Pp12 Nozzle CD

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Paper: ASAT-13-PP-12

13th International Conference on AEROSPACE SCIENCES & AVIATION TECHNOLOGY, ASAT- 13, May 26 28, 2009, E-Mail: asat@mtc.edu.eg Military Technical College, Kobry Elkobbah, Cairo, Egypt Tel : +(202) 24025292 24036138, Fax: +(202) 22621908

An Investigation on the Internal Flow in Simulated Solid Rocket Motor Chamber/Nozzle Configuration
M. Nasr*, A. M. Hegab** , W. A. El-Askary*, K. A. Yousif* Abstract: This paper describes numerical, analytical and experimental investigation of acoustic wave propagation in a simulated Solid Rocket Motor (SRM) chamber. The experimental study is carried out on a square-cylinder cross-sectional channel with two equally permeable sidewalls. An endwall disturbance is imparted using a moving piston located at the head end while the exit end of the channel is opened to the atmosphere. Moreover a convergent and convergent-divergent nozzle is changeable fixed at the exit end of the channel to study the behavior of the complex wave interactions mechanism at different nozzle areas. The unsteady, compressible, two dimensional Navier-Stokes equations in a laminar regime are numerically solved by predictor-corrector MacCormack scheme. Axial acoustic velocity field generated by the end wall disturbances interacts with steady sidewall injection to generate rotational flow field through the channel. As a result, a steady vorticity is generated at the sidewall and then is convected toward the centerline by the transverse component of the total velocity. Furthermore the time-independent, compressible NavierStokes equations with laminar effects are solved. An analytical solution for pure acoustic flow is derived from the reduced form of the full Navier-Stokes equations. The numerical and analytical solutions are compared with the experimental data. The comparisons show reasonable agreement between these three approaches. Moreover, the results show that, the geometry of the variable area parts has significant effect on the generated complex wave pattern inside the chamber. Keywords: Solid Rocket Motor chamber, Permeable sidewalls injection, Internal cavity, Solid rocket nozzle.

Nomenclature
bn C'o Cp Cv Et H H' K L
*

wave number speed of sound, m/s specific heat at constant pressure specific heat at constant volume total energy dimensionless channel half-height channel half-height, cm thermal conductivity dimensionless channel length

Mechanical Power Engineering Department, Faculty of Engineering, Menoufiya University, Shebin El-Kom, EGYPT ** Mechanical Power Engineering Department, Faculty of Engineering, Menoufiya University, Shebin El-Kom, EGYPT, Corresponding author: hegab2002us@yahoo.com, Tel. 012-7858517 1/24

Paper: ASAT-13-PP-12 L' M n P Patm P'o % P Pr Re T T'o t t'a u U'o v V'inj x X' y channel length, cm Mach number wave number index static pressure atmospheric pressure stagnation pressure, pa dimensionless pressure perturbation Prandtl number Reynolds Number temperature stagnation temperature, K time acoustic time, L'/ C'o axial speed reference axial speed, m/s transverse speed reference injection speed, m/s axial coordinate dimension axial coordinate transverse coordinate

Greek symbols ratio of specific heat aspect ratio transient axial velocity amplitude ' dynamic viscosity, pa.s density 'o stagnation density, Kg/m3 frequency Subscripts ' dimension quantities o stagnation value injection inj

1. Introduction
The current work is devoted to examine the time-dependent flow field in a porous channel with endwall disturbance to describe the effect of adding convergent and convergentdivergent area parts at the open end on the complex wave pattern inside the chamber. The presence of sidewall injection with traveling acoustic waves inside long slender square crosssectional channel can lead to a rotational flowfield that are decreed by the system geometry. These waves can, in turn, interact with the solid boundaries to generate acoustic and vortical wave resulting in complex flow patterns. In the hope of elucidating the nature of the resulting flowfield, an experimental investigation was conducted by Brown et al. [1]. They used nitrogen gas injection through uniformly sintered bronze plates inside a cylinder chamber. In their facility, acoustic waves are generated from an external rotary valve that controlled the flow exiting the chamber. Their results verified the accuracy of the analytical model suggested by Culick [2] for the mean flowfield and also provided substantial data for the resulting acoustic field. Independently of 2/24

Paper: ASAT-13-PP-12 the work given in [1], a novel investigation facility was built by Ma [3] to simulate similar flow conditions in a rectangular chamber. That experiment employed the sublimation of carbon dioxide, a process that resembled the combustion of propellant, in generating the CO2 gas to mimic the chamber's transpiring wall. Unfortunately, the work by Ma [3] had experimental difficulties in measuring acoustic pressure and velocity. The results indicated that the wave generator produced many non-harmonics waves with many higher fundamentals that make the interpretation of the complex mechanism is difficult. Moreover, in that experiment, Ma [3] didnt verify the occurrence of generating turbulence. Barron et al. [4] introduced an improvement to Ma's experiment by utilizing a Scotch-yoke mechanism to replace the slider-crank mechanism. The new mechanism led to higher pressure amplitudes, pure sinusoidal motion and resulted in acceptable validation between numerical and experimental results. Dunlap et al. [5] presented an experimental verification for the cold flow simulation of rocket chamber flow field based on Culick's analysis [2] of steady state flow. The results of Dunlap et al. [5] revealed that the inviscid flow field solution gives accurate results as long as the Reynolds number is sufficiently large to ensure that viscous effects are small compared with pressure gradients. Flandro [6] provided an early assessment of the importance of vorticity in acoustic boundary layer. He studied the impact of a small axial pressure gradient, varying harmonically in time, on the viscous process occurring adjacent to a surface from which a steady spatially uniform injection occurs. A linear equation for axial velocity contains a balance of convection, pressure gradient forces and viscous diffusion. The solution described a shear wave convecting away from the wall, with amplitude that is damped by viscous effects. Flandro [6] observed intense, transient vorticity in the boundary layer compared to the weaker steady vorticity associated with the inviscid, rotational Culick's solution [2]. A mathematical model formulated by Zhao et al. [7] is used to describe the initiation and evolution of intense unsteady vorticity in a low Mach number, weakly viscous internal flow sustained by mass addition through the side wall of a long narrow cylinder. The intense vorticity is formed at walls and is convected into the entire chamber by the steady radial velocity. The amplitude and the distribution of the vorticity are impacted by weak viscous and nonlinear effects. It was also demonstrated that there are parameter ranges of Mach number, Reynolds number and driving frequency for which vorticity is really confined to weakly viscous acoustic boundary layer, thin compared to the radius of the cylinder but larger than that obtained by Flandro [6]. Erickson et al. [8] presented analytical study concerning forced gas-dynamic oscillations in closed, constant diameter cylindrical ducts and ducts whose cross-section area varies axially. The objective of that work was to determine the effect of duct shape on resulting oscillations amplitude, wave form, harmonic content, and identify duct shapes that produce large amplitude oscillations for a given energy content. The results showed that the higher amplitude pressure oscillations can be forced in horn like shaped ducts as opposed to cylindrical ducts. Shocks like waveforms existed in constant diameter ducts are caused by the generation of higher harmonics through efficient non-linear coupling with the fundamental mode. In contrast, the non-linear coupling between the fundamental mode and its harmonic was weak in ducts whose cross sectional area varies axially (horn-like ducts). This induces much lower relative energy content in higher harmonics in that duct and hence decreases the excitation of harmonics. 3/24

Paper: ASAT-13-PP-12 The structure of turbulence in a channel flow with a fully transpired wall has been investigated experimentally by Deng et al. [9]. The aim of that experiment was to study the effect of the porous surface boundary conditions on the core flow development and flow structure in the channel. Air is sucked into the channel through the top honeycomb by high pressure direct drive blower. It was found that the boundary conditions on the porous surface are very important to the internal core flow evolution and flow pattern. For a course porous surface (1/4'' honeycomb), the mean flow differs significantly from classical Culick's solution [2] and computational results. However, with small pore size (1/8'' honeycomb), the mean velocity profiles are very close to laminar solution for a considerable downstream length and that profiles agree well with Culick [2], even though large turbulence intensity was observed. Hegab [10] presented numerical study that describes the transient flow dynamics generated in a SRM's chamber model with time-dependent mass injection. The main goal was to understand the heat transfer and temperature dynamics that accompany the co-existing acoustic and rotational velocity disturbances. Also the effect of adding variable duct to the open end on the internal flow-field was considered. The results showed that surprisingly large transient temperature gradient is presented at the sidewalls and the interior of the channel. Large gradients at the sidewall imply that there is an unexpectedly heat transfer which may influence the combustion zone above the burning propellant even though the fluid injection is isothermal. It is observed that the unsteady vorticity across the chamber is sensitive to small changes in flow dynamics and the maximum amplitude of the vorticity increases as the throat height is decreased. Also, the time history of pressure amplitude increases as throat height decreases. The influence of adding variable area portions to the open end of a circular constant-area tube on finite amplitude wave deformation and radiation, under resonant conditions was studied experimentally by Sileem and Nasr [11]. They concluded that, the radiated part of energy delivered by the piston to the atmosphere depends on the tube end configuration. The noticed difference between results of the open-end tube and that when variable area portions are added presumably attributed to the deviation of the natural frequency of the variable area cases from that of the open end tube case. The internal flow-field forming in the combustion chamber of SRM was analytically studied by Majdalani et al. [12]. A combined geometric configuration is considered in which a straight cylinder is connected to a tapered cone and the gases are injected perpendicularly to the surface. The selection of the injection velocity was followed from the experimental work of Brown et al. [1]. They concluded that the taper effect is more pronounced as the gases move away from the head end due to the increasing cross-sectional area. The mean flow approaches its asymptotic limit in sufficiently long cylinders. It may be worth mentioning that accurate matching of both numerical and analytical solutions requires that the motor parameters be chosen within specified limits and the corresponding criteria was shown to be practical. Similar experimental investigation to that of [12] used for studying the influence of solid propellant inclination angle through small-scale cold flow simulation was presented by Nguyen et al. [13]. The experiments were conducted on a cold gas experimental setup. They concluded that the sidewall injection angle has significant effect on the internal flow field and vortex shedding in SRM. For small inclination angles the whole interactions between shear layer and vortex shedding were decreased and hence decreased the wall vorticity, weaken pressure wave intensity and oscillation levels. On the contrary, the larger angles enhanced the influence of wall vortex shedding at the rear end of the chamber, which lead to the increase of pressure fluctuation levels. 4/24

Paper: ASAT-13-PP-12 Hegab and Nasr [14] studied experimentally and numerically the propagation of acoustic waves (which is generated by oscillating piston) in a long, narrow chamber with endwall disturbance. Their study includes also, an analytical solution to the two following cases. The first case is in which straight duct with endwall disturbance at the head end and without sidewall mass injection. The second case, duct contributes steady sidewall mass injection from permeable walls and endwall disturbance. They illustrated a reasonable comparison between the experimental, analytical, and the computational results using the two-four explicit predictor-corrector MaCcormack scheme. From the mentioned review it is noticed that, the effect of exit geometry on the acoustic flowfield generated in SRM cavity with/without sidewall mass injection did not take the sufficient attention. Therefore, the present work focuses on experimental and theoretical studies for the unsteady flow in a simulated SRM interior cavity. The cavity is represented by square-cylinder chamber with permeable sidewalls. Moreover, convergent and convergentdivergent nozzles are changeable installed at the exit end of the channel. The numerical solution of the laminar, two dimensional, compressible and steady/unsteady Navier-Stokes equations for the same geometry are considered to validate the experimental work. Also the analytical solution of the reduced form of Navier-Stokes equations is considered.

2. Experimental Setup
The cavity of SRM is represented by square-cylinder chamber of axial length 44 cm and cross section height of 2.5 cm. The general arrangement of the experimental set-up layout is shown in Fig.1. It consists of the following parts: an oscillating piston [2.5 cm diameter and 1.9 cm stroke] driven by an electric AC motor of YC90S-2 type, 2900 rpm, 1 HP, single phase electrical input and suitable for variable speed through a pulley-belt system. The number of revolution of AC motor can be changed (from 3610 rpm to 4774 rpm) by voltage regulator. The variable speed piston oscillates to generate the acoustic disturbance at one end of the channel and the other end is open to atmosphere. The square cross-sectional area channel has permeable sidewalls to inject a guided-steady state flow into the channel. The injected air flows through two similar hoses which connected to two similar injection ducted guided-blade ducts to obtain uniform injection steady velocity at both sides of the channel. Convergent and convergent-divergent nozzles are changeably added to the open end of the square crosssectional area channel. The detailed dimensions of those parts are given in Table (1). Wall pressure time histories are recorded at two locations along the channel, where the lower one is 5.5 cm apart from the Top Dead Center (TDC) and the upper one is at 19.5 cm apart from square cross-sectional area channel end. These pressure histories are measured using Capacitance Pressure Transducer [model SA, Data Instrument, Action MA01720, USA 0-100 psi, accuracy 1%] connected to Data Acquisition System (DAS) and desktop computer through a Labview software. The data are recorded in files and plotted as will be shown later. The uncertainties of the measured pressure shown here are estimated according to the procedures given by Coleman et al. [15] found to be within 2.5% for a confidence interval of 95.45 %. The reading error of U-tube manometer which is used to measure the injected flow rate by an orifice meter is 0.25 mm and the total uncertainty of the measured flow rate is equal to 1.412 106 .

5/24

Paper: ASAT-13-PP-12

(1) (2)

(3)

(4)

(18) (9) (7)


(11)

~
(10)

(8)

(6) (5)

(13)

T.D.C
(14)

B.D.C
(17) (16)

(15)

( 1) 2

(5)
head end

(8)

2H
X

2 H t h

2 H ex

L c

L cd

1- Air from atmosphere 2- Reciprocating compressor 3- Air reservoir (volume 500 lit.) 4- Orifice meter 5-Two perforated plates 6- Injection air flow through duct with guide blades 7- Square channel (test duct) 8- Electrical power supply 9- Controlled voltage circuit and constant voltage power supply device

10-Distribution board 11- Pressure measuring holes connected by capacitance pressure transducer 12 To DAS 13- Desktop computer 14- Oscillating piston 15- Pulley-belt system 16- AC motor 17- Voltage regulator 18- U-tube manometer

Fig.1 The experimental test-rig components Table (1) Detailed dimensions of straight, convergent, and convergent-divergent area channels. Length Exit height, Throat height, Part type (cm) 2H'ex (cm) 2H'th (cm) Square Duct Constant area 44 2.5 ----------Convergent Variable area 5 0.65,0.95,1.27 ----------ConvergentVariable area 14 3,5.78,8.725 0.65,0.95,1.27

6/24

Paper: ASAT-13-PP-12

3. Mathematical Model
The mathematical model is based on solving the unsteady, two dimensional parabolized Navier-Stokes equations describing both dynamics and acoustic in a perfect gas within a square-cylinder channel. The gas is initially stationary with a reference-state defined by Po (the total pressure), o (the total stagnation density) and T o (the total temperature). The

associated sound speed is found to be C o = ( Po / o ) , where prime quantities refer to


dimensional values. The characteristic length scales for the axial and transverse variables are chosen to be the axial length of the channel L and the half-height of the cross section H , respectively. Time is nondimensionalized with respect to the axial acoustic time t a = L / C o . The aspect ratio is defined as the ratio of the channel length to the half-height of the channel, = L / H .The characteristic axial velocity U o is related to the characteristic side wall injection velocity of the fluid V inj through the global mass conservation. The characteristic axial flow Mach number, Prandtl number and flow Reynolds number are defined respectively, as follows: M = U o /C o , Pr = C po /k , Re = oU o L /
The governing flow equation can be written in its dimensionless form as follows;

Q F + + =0 t x y
Where Q, E and F are column vectors given by [16]

(1)

u u u + p Q= v , = uv t

M v uv u y / Re , F = v + p /( ) y u t + ( 1)up v [ t + ( 1) p ] Re . Pr

The dimensionless form of the equation of state for a perfect gas is;

p = T
The non dimensional form of the total energy is represented by

(2)

v ( 1) [u + ( ) ] t = Cv + 2

(3)

Where, Cv and are the specific heat at constant volume and the specific heat ratio, respectively. The velocity and temperature gradients read;

uy =

u y

,T

(4)

7/24

Paper: ASAT-13-PP-12 In general >>1, M<1, Re>>1 and Pr=O(1) In the present work, two solution techniques including numerical as well as analytical methods are used. They will be discussed in the following subsection.

3.1. Numerical Method


An accurate flow-field time history in SRM chamber can be obtained using finite-difference scheme which shows the evolution of the flow variables in the axial and transverse directions after many acoustic wave cycles. The present study employs higher-order accuracy difference equations to minimize the impact of numerical diffusion which was found to affect the results obtained from the second order explicit MacCormack code, see Hegab [17]. Near the boundaries, the second order explicit predictor-corrector scheme, developed by MacCormack [18] is used to descretize the two-dimensional, unsteady, compressible, Navier-Stokes equations. At the interior points, the Navier-Stokes equations are descretized using the TwoFour explicit, predictor corrector scheme, developed by Gottlieb and Turkel [19], which is a fourth-order variant of the fully explicit MacCormack scheme. This method, applied by Kirkkopru et al. [20], is highly-phase accurate and therefore suitable for describing many wave cycles and wave interaction problems. Hegab [21] used the same numerical method in similar problems with good numerical predictions.

3.1.a Boundary Conditions


There are three kinds of boundary conditions used in this study. The first, case (1) in which the air is injected uniformly across the sidewalls of the channel and the head end is closed while the other free end is open to atmosphere. Also, variable area nozzles are changeably added at the open end. At the head end, x=0 the no slip boundary conditions are applied, u= v=0. While, at the duct exit, the pressure is taken as an atmospheric pressure and the other parameters are extrapolated using zero gradient at exit. At the sidewalls, the no slip condition is applied for the streamwise component of velocity, but the crosswise velocity is taken to be the injection velocity V'inj. The second, case (2) of the boundary conditions is concerned with studying the internal flow field when only acoustic waves are generated at the closed end while the other free end is open to atmosphere. Also variable area nozzles are added to the channel free end. The same boundary conditions for the case (1) are applied except that; the no slip condition is applied also to the velocity components at sidewalls, i.e. u=v=0. At the head end, x=0, the streamwise velocity is considered in the form u= sin (t). In the third, case (3) of boundary conditions the air is injected uniformly across the sidewall of the chamber and acoustic waves are generated at the head end while the other free end is open to atmosphere. The variable area nozzles are also included at the free end. The same boundary conditions for the case (1) are applied except that; at the head end (x=0), u= sin (t). At the sidewalls, v'(X', 0) =V'inj and v'(X', 2H') =-V'inj, (v=v'/V'inj, V'inj=0.1273 m/s)

3.2. Analytical Solution 3.2.a Analytical Solution for Case (1)


In this case a steady uniform injection from the sidewalls of square-cylinder channel is considered. Culick [2] and Hegab [10] showed that the axial velocity distribution is independent of viscosity when Reynolds number is sufficiently large. Culick [2] derived the following equations for axial, transverse velocities and static pressure:

u y = cos[( ) (1 )] Uo 2 H
8/24

(5-a)

Paper: ASAT-13-PP-12
v y = sin[( ) (1 )] Vinj 2 H Po P = ( / L )2 Po P atm

(5-b) (5-c)

Where Uo is the center line dimensionless velocity at the channel exit, and Po is the dimensionless stagnation pressure at the channel head end.

3.2.b Analytical Solution for Case (2)


In the case (2), endwall disturbance is fitted at the duct head end while the other end is open to atmosphere. The analytical approach is based on the reduced form of the Navier-Stokes equations using asymptotic techniques [9, 14 and 20]. The final solution for pressure and velocity is as follows: The non resonance acoustic solution for axial velocity is: % u (x , t ) = sin t + 2
n =0

( sin t sin b n t ) sin b n x for bn* b 2 b n2


2 n

(6)

Where bn is the wave number, b n = (n + 1/ 2) with n as the wave number index, is the forced frequency and bn* is the wave number at resonance conditions. The perturbation % pressure p ( x , t ) can be obtained using the following equation [20]:

% u% ( x , t ) 1 p (x ,t ) = t x

(7)

The boundary condition at the duct head end, x=1 reads p=1. Integrating equation (7) yields the pressure perturbation in the form [20]:

% p (t , x ) = (x 1)cos t

2 (cos t cosbnt )cosbn x 2 n =0 b n

(8)

The solution for resonance case can be also found in [20] as:

1 % u (x , t ) = sin t { sin(b n *t ) + t cos(b n *t )}sin(b n *x ) bn *


% p (x , t ) = cos(t ) (x 1) 2 cos b *t b n *t sin b *t cos b * x
n n n

(9)

Where b n * = and represents the wave amplitude

4. Results and Discussions


The present results are classified into three categories. The first category of the results (case (1)) deals with the steady state internal flow generated by a steady mass addition from the two sidewalls of the square-cylinder channel (nozzless duct). Moreover, the effect of adding convergent and convergent-divergent area nozzles to the end of the duct is considered. The second category of the results (case (2)) considers the same channel with endwall disturbance 9/24

Paper: ASAT-13-PP-12 at the head end, while variable area nozzles are added changeably at the channel exit. The last category (case (3)) deals with the channel as in the second category with the presence of uniform injected air from permeable sidewalls. The axial pressure distribution along the chamber wall is shown Fig. 2 for experimental, numerical and analytical solutions as in case (1). In such case, air is injected uniformly across the sidewalls of the square chamber with V inj = 0.1273m / s and the head end is closed while
the other free end is open to atmosphere (nozzless duct). It is noticed that, the static pressure reaches its maximum value at the closed duct head end at ( X / L = 0 ) and gradually decreases toward the channel exit. The comparison between the numerical and analytical approaches with the experimental data shows reasonable agreement. In Fig. 3, the axial velocity distribution across the channel height from the present numerical code, the analytical solution and the results obtained using the commercial code Fluent [22] at different axial locations is shown. The development of the velocity profiles is attributed to the mass addition from the walls as axial distance increases. It is noted that the velocity profiles are found to be very close to the laminar velocity profiles and symmetrical about the centerline. Consistent results for Fig. 3 have been shown also in [10]. Figure 4 shows transverse velocity at different axial locations for the same conditions given in Fig. 3. The numerical and analytical results are almost coincides in the all domain, while small deviation occurs near the duct inlet and exit. This deviation is attributed to the increase in the axial velocity due to the mass accumulation at aft end and in turn makes the flow to be convected towards downstream as soon as it is injected from the walls. However, the analytical solution is proved to be the best choice to simulate such flowfield case when the air is injected uniformly along the sidewall, like the gasification of propellant in SRMs.
1.00050 1.00045 1.00040 1.00035

Numerical 1 Analytical

Experimental data

P ' / P 'a t m

1.00030 1.00025 1.00020 1.00015 1.00010 1.00005 1.00000

0.1

0.2

0.3

0.4

0.5 X' / L'

0.6

0.7

0.8

0.9

Fig. 2 Static Pressure distribution along the duct wall for nozzless duct without acoustic (case (1)).[1-persent code]

10/24

Paper: ASAT-13-PP-12
1.0 0.9 0.8 0.7

Numerical 1 Numerical 2 Analytical

Y ' /2 H '

0.6 0.5 0.4 0.3 0.2 0.1 0.0

U' (m/s)

Fig. 3 Transverse axial velocity distribution at different axial locations for nozzless duct without acoustic (case (1)).[1-persent code, 2-Fluent code]
1.0

X=0.25
0.9 0.8 0.7

X=0.75 X=0.95 Analytical

X=0.5

Y ' /2 H '

0.6 0.5 0.4 0.3 0.2 0.1 0.0 -0.15 0.0

-0.1

-0.05

v'(m/s)

0.05

0.1

0.15

Fig. 4 Transverse velocity distribution for nozzless channel without acoustic (case (1)).

The effect of adding convergent area nozzle on the static pressure distribution along the wall of the channel for case (1) is shown in Fig. 5. The convergent area nozzle has total height at exit of 2H ex = 0.65cm and its converging length about 5 cm. It is appeared that the static pressure decreases slowly through the channel and then it decreases with strong gradient through the nozzle due to the flow acceleration in the nozzle. The over-prediction of the experimental results with numerical may be related to uncertainty of the experimental data.

11/24

Paper: ASAT-13-PP-12

L'

L'c

H' H'e x
1.005

Experimental Numerical 1
1.004

P ' / P 'a t m

1.003

1.002

1.001

1.000

0.0

0.2

0.4

X'/L'

0.6

0.8

1.0

1.1136

Fig. 5 Static Pressure distribution along the wall of duct ended by convergent nozzle, 2H'in=2.5 cm, 2H'ex=0.65 cm, L'c=5 cm and V'inj=0.1273 m/s, (case (1)). The second set of the results for case (2) discusses the internal flowfield due to endwall disturbance using reciprocating piston at the head end of the channel along with and without a convergent nozzle added at the end of the channel. The main purpose of this set is to describe the effects of forced frequency, axial location and the nozzle geometry on the generated acoustic field inside the interior cavity. In Figs. 6 and 7, a comparison between the experimental and numerical results for nozzless duct at the same dimensional flow characteristics as in Fig. 5 is presented. These Figures demonstrate the pressure time history at two dimensionless distances from the head end or to be more precise from the top dead center of the piston ( X / L = 0.125 and 0.557 ) and constant dimensionless forced frequency =0.65. It is observed that, there is small deviation in the static pressure mean value and the phase angle between the experimental and numerical results. Moreover, higher harmonic oscillations are seen in the experimental data which represents the eigenfunction mode contributions. The reasons behind these deficiencies may be related to the sudden change of cross-sectional area from circular duct (in which the piton moves) at the TDC to the square cross-sectional channel. Moreover, using slider-crank shaft mechanism cant produce a pure sinusoidal piston motion. Since the piston movement is not purely sinusoidal, many harmonics are introduced into the acoustic velocity and pressure fields as discussed by Barron et. al [4]. Furthermore, it is found that, the amplitude of the measured static pressure oscillation decreases as the axial distance increases downstream. That may be attributed to the wave attenuation due to the boundary layer development on the duct sidewalls and acoustic streaming.

12/24

Paper: ASAT-13-PP-12

R c

L c

u = sin t
Y X

head end

2H L

1.008

N umerical 1 Experimenta l
1.004

P ' / P 'a t m

1.000

0.996

0.992

20

40

60

80

100

Fig. 6 Pressure-time history for nozzless channel fitted by forced oscillation, =0.65, =0.045 and X'/L'=0.125 (case (2)).
1.008

N umerical 1 Experimental
1.004

P ' / P 'a t m

1.000

0.996

0.992

20

40

60

80

100

Fig. 7 Pressure-time history for nozzless channel fitted by forced oscillation, = 0.65, =0.045 and X'/L'=0.557 (case (2)).

13/24

Paper: ASAT-13-PP-12 Figure 8 represents a comparison between analytical and numerical solutions for the pressure time history at =1.1, =0.1, X / L = 0.125 and =20. There are small deviations in both phase and amplitude between the numerical and analytical results. This deviation may be related to the assumption with the analytical solution that considers an incompressible, inviscid flow and linear system. Similar trend was found by Hegab [10]. During one complete cycle of the piston, the acoustic wave completes two round trips across the channel length as discussed by Sileem and Nasr [11]. A comparison of the analytical and numerical results near resonance frequency for pressure time history is presented in Fig. 9 at =1.55, =0.1, X / L = 0.557 and =20. For t10, the results show good agreement in amplitude and phase
1.060

Numerical 1
1.040

Analytical

1.020

P ' / P 'a t m

1.000

0.980

0.960

0.940

20

40

60

80

100

Fig. 8 Pressure-time history for nozzless channel fitted by forced oscillation, =1.1, X'/L'=0.125, M=0.1 and =0.1 (case (2)).
1.5 1.4 1.3 1.2

Numerical 1

Analytical

P ' / P 'a t m

1.1 1.0 0.9 0.8 0.7 0.6 0.5

10

20

30

40

50

Fig. 9 Pressure-time history for nozzless channel fitted by forced oscillation, =1.55, X'/L'=0.557, M=0.1 and =0.1 ((case (2)).

14/24

Paper: ASAT-13-PP-12 between numerical and analytical approaches. While for t>10 the analytical solution shows linear growth with time rather than beats with the numerical solution. Similar results were previously noticed by Hegab [21], who discussed this phenomenon and gave interpretation to this trend as the nonlinearity effect with the numerical solution. The effect of forced frequency on the axial velocity time history is introduced in Figs. 10 and 11. Data are recorded at X / L = 0.125 and =0.52 and 0.65, respectively. It is observed that, increasing the forced frequency leads to an increase of the axial velocity amplitude with the same phase angle.
0.3

Numerical 1
0.2

Analytical

0.1

u ' / U 'o

-0.1

-0.2

-0.3

20

40

60

80

100

Fig. 10 Axial velocity time-history for nozzless channel fitted by forced oscillation, =0.52, X'/L'=0.125 and =0.1 (case (2)).
0.3

Numerical 1
0.2

Analytical

0.1

u ' / U 'o

-0.1

-0.2

-0.3

20

40

60

80

100

Fig. 11 Axial velocity time-history for nozzless channel fitted by forced oscillation, =0.65, X'/L'=0.125 and =0.1 (case (2)).

15/24

Paper: ASAT-13-PP-12 Figures 12 and 13, show the acoustic axial velocity across the channel at x=0.25 and 0.5 for =1.5, M=0.1 and =0.1 at four times t=10, 20, 40 and 78. One can observe the presence of the acoustic boundary layer development at the same location with different boundary layer thickness. Rotational flow is considered to be confined in that region. The explanation is related to the interaction between the duct sidewalls and the acoustic wave motion. Baum [23] noticed an overshoot of the axial velocity at the edge of acoustic boundary layer and showed that, acoustic boundary layer thickness and axial velocity overshoot at its edge are strongly influenced by mean flow Reynolds number and frequency of oscillations.
1.00

0.90

Y '/2 H '

0.80

0.70

0.60

t=10 t=20 t=40 t=78

0.50

-1

-0.5

u' / U'o

0.5

1.5

Fig. 12 Transverse axial velocity distribution for nozzless channel fitted by forced oscillation, =1.5, X'/L'=0.25 and =0.1 and (case (2)).
1.00

0.90

Y ' /2 H '

0.80

0.70

0.60

t=10 t=20 t=40 t=78


-1 -0.5 0 0.5 1 1.5 2

0.50

u' / U'o

Fig. 13 Transverse axial velocity distribution for nozzless channel fitted by forced oscillation, =1.5, X'/L'=0.5 and =0.1 (case (2)).

16/24

Paper: ASAT-13-PP-12 The effect of installed convergent and convergent-divergent nozzles at the end of the channel on the complex wave interaction mechanism inside the chamber is illustrated experimentally in Figs. 14 and 15, respectively. The angular frequency is taken to be =0.52, and the convergent nozzle length and total exit height are Lc = 5cm , 2H ex = 0.65cm , while the convergent-divergent nozzle length, total throat and exit heights are Lcd = 14cm , 2H th = 0.65cm and 2H ex = 3cm , respectively. The exit height for the convergent area nozzle is equal to the throat height of the convergent-divergent area nozzle with an equal converging length. It is shown that, the static pressure amplitude for channel ended with nozzles (at x=0.125) has higher amplitude with deviation in the mean value compared to nozzless duct. It may be attributed to the decreasing of the flow area at the nozzle exit which leads to a reduction in transmitted energy waves from nozzle exit and hence increases the interactions between incident and reflected waves from boundaries which in turn increases the pressure amplitude. The nozzle and nozzless results show higher harmonic oscillations as discussed by Barron et. al [4].
1.008

Nozzless channel Channel ended by convergent nozzle


1.004

P ' / P 'a t m

1.000

0.996

0.992

20

40

60

80

100

Fig. 14 Pressure-time history for channel fitted by forced oscillation ended by convergent nozzle,(2H'ex=0.65 cm) and nozzless channel at =0.045, =0.52 and X'/L'=0.125 (case (2)).

17/24

Paper: ASAT-13-PP-12
1.008

Nozzless channel Channel ended by convergent-divergent nozzle


1.004

P ' / P 'a t m

1.000

0.996

0.992
0 20 40

60

80

100

Fig. 15 Pressure-time history for channel fitted by forced oscillation ended by convergent-divergent, (2H'th=0.65 cm, 2H'ex=3 cm)and nozzles channel at X'/L'=0.125, =0.52 and =0.045 (case (2)).

Moreover, the significant difference between the convergent and convergent-divergent nozzle results is presumably related to the deviation of the natural frequency producing by adding variable area portions at the end of the duct. Also, it is noticed that, the effect of adding convergent-divergent nozzle results in higher amplitude in static pressure but still lower amplitude compared with that of convergent nozzle. Sileem and Nasr [11] and Hegab [10] have noticed similar trend in their results. The effect of adding convergent and convergent-divergent nozzle at the open end of the channel with sidewall mass addition is shown in Fig. 16 for case (3). The sidewall injection velocity is V inj = 0.1273m / s and endwall disturbance at the head end with =0.65. It is shown that, the steady sidewall injection reduces the pressure amplitude significantly. (Compare Figs. 15 and 16). It may be explained that, the flow is injected through porous surface on the duct sidewalls across two parallel perforated plates which act as jet velocity through the generated flow field and hence attenuate the acoustic energy especially when that energy is nearly small and the effect of exit from the nozzle doesnt play an explicit role in the pressure amplitude.

18/24

Paper: ASAT-13-PP-12
1.008 Nozzless channel Channel ended by convergent nozzle Channel ended by convergent-divergent nozzle 1.004

P ' / P 'a t m

1.000

0.996

0.992

20

40

60

80

100

Fig. 16 Pressure-time history for nozzless channel, channel ended by convergent nozzle and convergent-divergent at forced frequency =0.65, =0.045 and X'/L'=0.125 (case (3)).

The effect of the exit height on the acoustic flowfield generated in the duct ended by convergent nozzle (without air flow injection) is shown in Fig. 17 for case (2) at X / L ' = 0.125 , = 0.52 and different exit heights; 2H ex = 1.27, 0.95 and 0.65cm . It is shown from this figure that, with deceasing the exit height of the convergent nozzle, the pressure amplitude increases. It is clearly seen by decreasing the exit height causes an increase in the amount of reflected acoustic energy which in turn interacts with the incident waves from the piston and hence enhances the pressure amplitude. Figure 18 shows the effect of throat height on the pressure time history for duct ended with convergent-divergent nozzle. It is concluded that, at the same forced frequency and axial location, for no air flow injection; decreasing the throat height increases the pressure amplitude but still smaller than the effect of adding a convergent nozzle with the same throat heights. The power spectrum density (PSD) is calculated from the experimental data for case (2) and is shown in Fig. 19 for nozzless channel and without injection. It is shown that, the gas oscillates at the same frequency of the piston (79.6 HZ). Higher harmonics of very small energy content could be noticed before and after the main frequency. The effect of adding convergent nozzle by different exit heights on the energy content in PSD is shown in Fig. 20. It is concluded that, decreasing the exit height of the convergent nozzle increases the amount of reflected waves which in turn increases the wave's interactions and energy content. Moreover, higher harmonics increases as the exit height of the convergent nozzle decreases. This is an indication of small wave deformation. This confirms the aforementioned understanding of the mechanism of wave amplitude increasing, which is mainly due to off-resonant situation caused by the exit area reduction.

19/24

Paper: ASAT-13-PP-12
1.008

2H'ex=1.27 cm 2H'ex=0.95 cm 2H'ex=0.65 cm


1.004

P ' / P 'a t m

1.000

0.996

0.992

20

40

60

80

100

Fig. 17 Pressure-time history for the channel ended by convergent nozzle at different exit heights at forced frequency =0.52, =0.045 and X'/L'=0.125 (case (2)).

1.008

2H'th=1.27 cm 2H'th=0.95 cm 2H'th=0.65 cm

1.004

P ' / P 'a t m

1.000

0.996

0.992

20

40

60

80

100

Fig. 18 Pressure-time history for the channel ended by convergent-divergent nozzle at different throat heights at forced frequency =0.65, =0.045 and X'/L'=0.125 (case (2)).

20/24

Paper: ASAT-13-PP-12
0.0002

0.00016

P SD

0.00012

8E-005

4E-005

40

80

120

160

200

Fig. 19 Power spectrum density for nozzless channel at X'/L'=0.125 (case (2)).
0.0002

2H'ex=1.27 cm
0.00016

2H'ex=0.95 cm 2H'ex=0.65 cm

0.00012

PSD
8E-005 4E-005

40

80

120

160

200

Fig. 20 Power spectrum density for channel ended by convergent nozzle at X'/L'=0.125 (case (2)).

PSD for channel ended by convergent-divergent nozzle is shown in Fig. 21. It is obvious that, installing convergent-divergent nozzle at the channel end increases the energy content of the wave. Moreover, higher harmonics are clearly seen in this figure. It may be related to energy dissipation and transfer from the fundamental mode to its higher harmonics. Also the frequency deviation for the three throat heights may be result from the difference of the natural frequency between the nozzless channel and channel ended with convergent-divergent nozzle. Moreover, decreasing the throat height increases the PSD at the same frequency. The maximum and minimum pressure amplitudes for nozzless channel, channel ended by convergent nozzles and convergent-divergent nozzles are shown in tables (2), (3) and (4), respectively. 21/24

Paper: ASAT-13-PP-12
0.0002

2H'th=1.27 cm 2H'th =0.95 cm


0.00016

2H'th=0.65 cm

P S D

0.00012

8E-005

4E-005

40

80

120

160

200

Fig. 21 Power spectrum density for channel ended by convergent-divergent nozzle at X'/L'=0.125 (case (2)). Table (2) The maximum and minimum pressure amplitudes for nozzless channel. Pmin Pmax

0.52 0.63 0.65

0.9955 0.99515 0.99405

1.0029 1.003 1.0032

Table (3) The maximum and minimum pressure amplitudes for channel ended by convergent nozzle.

2H'ex=1.27 cm 0.52 0.63 0.65 Pmin 0.99732 0.9953 0.99517 Pmax 1.0026 1.00275 1.0028

2H'ex=0.95 cm Pmin 0.99612 0.99525 0.99745 Pmax 1.0038 1.0039 1.0033

2H'ex=0.65 cm Pmin 0.99512 0.995 0.99472 Pmax 1.00425 1.0049 1.00527

Table (4) The maximum and minimum pressure amplitudes for channel ended by convergent-divergent nozzle. 2H'th=0.65 cm 2H'th=0.95 cm 2H'th=1.27 cm and and and 2H'ex=3 cm 2H'ex=8.725 cm 2H'ex=5.78 cm

0.52 0.63 0.65

Pmin 0.99599 0.99633 0.99521

Pmax 1.00388 1.00498 1.0052

Pmin 0.9951 0.9952 0.9956 22/24

Pmax 1.00382 1.0041 1.0042

Pmin 0.99737 0.9955 0.99618

Pmax 1.0027 1.0038 1.00395

Paper: ASAT-13-PP-12

5. Conclusions
The effect of adding convergent and convergent-divergent nozzles at the end of squarecylinder channel on the internal flow field is studied experimentally, analytically and numerically. The flow field in the channel is generated by either steady mass addition from sidewalls and/or endwall disturbance. The following statements could be drawn from the current study: 1. For the considered injection mass rate, the whole flow is found to be subsonic laminar. 2. The flow is highly rotational near walls and it then weakens until reaching the core region as demonstrated in case (2). 3. The internal acoustic flowfield is highly affected by some parameters such as forced frequency, boundary conditions treatment and exit area geometry. 4. Adding a convergent nozzle at the end of the channel leads to partially transmitted acoustic waves at the duct exit and the remained are reflected from the solid walls. The latter increases when the exit height decreases and hence increasing the wave amplitude. Finally, the effect of convergent-divergent nozzle existence at the end of the channel shows wave interaction mechanism, which is completely different than that with either nozzless or convergent nozzle existence. Moreover, the flow injected through the porous sidewall attenuates the generated acoustic field throughout the channel. These acoustic waves interact with the steady sidewall injection to generate vorticity across the chamber. The generated rotational flow may impact the burning rate in the real solid rocket motor propellant.

6. Acknowledgment
This work is supported by the Science and Technology Development Fund (STDF) through the project ID-108.

References:
[1] [2] [3] [4] [5] [6] [7] [8] Brown, R.S., Blackner, A.M., Willoughby, P.G., and Dunlap, R., "Coupling between Acoustic Velocity Oscillations and Solid Propellant Combustion", J. of Propulsion and Power, Vol. 2, No.5, 1986, pp. 428-438. Culick, F.E.C., "Rotational Axisymmetric Mean Flow and Damping of Acoustic Waves in a Solid Propellant Rocket", AIAA Journal, Vol. 4, No. 8, 1996, pp.1462-1464. Ma, Y., "A Simulation of the Flow Near a Burning Propellant in a Solid Propellant Rocket Motor", Ph D. Thesis, University of Utah, Salt Lake City, USA, 1990. Barron, T., Van Moorhem, W.K, and Majdalani, J., "A Novel Investigation of the Oscillatory Field over a Transpiring Surface", J. of Sound and Vibration, Vol. 235, No. 2, 2000, pp. 281-297. Dunlap, R., Willoughby, P.G, and Hermsen, R.W., "Flow-Field in the Combustion Chamber of a Solid Propellant Rocket Motor", AIAA journal, Vol. 12, No. 10, 1974, pp.1440-1442. Flandro, G.A., "Solid Propellant Acoustic Admittance Corrections", J. of Sound and Vibration, Vol. 36, No. 3, 1974, pp. 297-312. Zhao, Q., Kassoy, D.R, and Kirkkorpru, K., "Acoustically Generated Vorticity in an Internal Flow", J. of Fluid Mech., Vol. 413, pp. 247-285. Erickson, R., Markopoulos, N., and Zinn, B.T., "Finite Amplitude Acoustic Waves in Variable Area Ducts", AIAA paper 2001-1099, 39th Aerospace Sciences Meeting and Exhibit, Jan 8-11, Reno, NV., USA.

23/24

Paper: ASAT-13-PP-12 [9] Deng, Z., Addrian, R.J, and Tomkinson, C.D.," Structure of Turbulence in Channel Flow with a Fully Transpired Wall", AIAA paper 2001-1019, 39th Aerospace

Sciences Meeting and Exhibit, Jan. 8-11, Reno, NV, USA.


[10] Hegab, A.M., "Study of Acoustic Phenomena in a Solid Rocket Engines", Ph.D. Thesis, Mech. Power Engineering Dept., Menoufiya Univ., Egypt, carried out at Center for Combustion Research, University of Colorado at Boulder, USA ,1998. [11] Sileem, A.A., and Nasr, M., "An Experimental Study of Finite Amplitude Oscillations in Ducts: The Effect of Adding Variable Area Part to the Open End of Constant-Area Resonant Tube", Alexandria Engineering Journal, Vol.42, No. 4, 2003, pp. 379-409. [12] Majdalani, J., Sams IV O.C., and Saad, T., "Mean Flow Approximations for Solid Rocket Motors with Tapered Walls", J. of Propulsion and Power, Vol. 23, No2. , 2007,

pp. 445-456.
[13] Nguyen, C.C., Plourd, F., and Doam, S.K., "Analysis of Injecting Wall Inclination on Segmented Solid Rocket Motor Instability", J. of Propulsion and Power, Vol. 24. No. 2, 2008, pp. 213-223. [14] Hegab, A.M., and Nasr, M., "Unsteady Vorticity Fields in a Long Narrow Channel with Endwall Disturbances and Sidewall Injection", EG-220, 9th International Congress of

Fluid Dynamics & Propulsion, Dec. 18-21, 2008, Alexandria, Egypt.


[15] Coleman, W.H., and Glenn S.W., "Engineering Application of Experimental Uncertainty Analysis", AIAA J, Vol. 33, No. 10, 1995, pp. 1888-1896. [16] Anderson, J.D., "Computational Fluid Dynamics", McGraw-Hill, Inc, New York, NY, USA, 1995. [17] Hegab, A.M., " Vorticity Generation and Acoustic Resonance of Simulated Solid Rocket Motor Chamber with Transient Sidewall Mass Injection", Proceedings of the 9th International Congress of Fluid Dynamics & Propulsion, ICFDP9-265 December 18-21, 2008, Alexandria, Egypt. [18] MacCormack, R.W., "The Effect of Viscosity in Hypervelocity Impact Crating", AIAA Hypervelocity Impact Conference, Cincinnati, Ohio, April, 1969, No.69-354. [19] Gottlieb, D., and Turkel, E., "Dissipative Two-Four Methods for Time-Dependent Problems", Int. J. Mathematics and Computation, Vol. 30, No. 136, 1976, pp. 703-

723.
[20] Kirkkopru, K., Kassoy, D.R., and Zhao, Q., "Unsteady Vorticity Generation and Evaluation in a Model of Solid Rocket Motor", J. of Propulsion and Power, Vol. 12, No.4, 1996, pp. 646-654. [21] Hegab, A.M., "Vorticity Generation and Acoustic Resonance of Simulated Solid Rocket Motor Chamber with High Wave Number Wall Injection", J. of Computers & Fluids,

Vol. 38, No. 6, 2009, pp. 1258-1269.


[22] Fluent-6.3.26. User's Manual, (2006), Fluent Inc., Lebanon, USA [23] Baum, J.D., Investigation of Flow Turning Phenomenon Effects of Frequency and Blowing Rate, AIAA paper 1989-0297, 27th Aerospace Sciences Meeting and Exhibit, Jan. 9-12, Reno, NV., USA.

24/24

You might also like